• Aucun résultat trouvé

Phenolic acids interactions with clay minerals: A spotlight on the adsorption mechanisms of Gallic Acid onto montmorillonite

N/A
N/A
Protected

Academic year: 2021

Partager "Phenolic acids interactions with clay minerals: A spotlight on the adsorption mechanisms of Gallic Acid onto montmorillonite"

Copied!
37
0
0

Texte intégral

(1)

HAL Id: insu-02164418

https://hal-insu.archives-ouvertes.fr/insu-02164418

Submitted on 11 Sep 2019

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Phenolic acids interactions with clay minerals: A

spotlight on the adsorption mechanisms of Gallic Acid

onto montmorillonite

Adoum Mahamat Ahmat, Thomas Thiebault, Régis Guégan

To cite this version:

(2)

Phenolic acids interactions with clay minerals: a spotlight on the adsorption

mechanisms of Gallic Acid onto Montmorillonite

Adoum Mahamat Ahmat(a,b)1, Thomas Thiebault(c)

, Régis Guégan(b,d) 2

(a) Institut Mines-Telecom Lille-Douai, Department of Civil Engineering and Environment. 764, Boulevard Lahure, 59508 Douai, France.

(b) Institut des Sciences de la Terre d’Orléans ISTO, UMR 7327, University of Orléans, 1A Rue de la Férollerie, 45500 Orléans La source, France.

(c) EPHE, PSL University, UMR 7619 METIS (SU, CNRS, EPHE), 4 place Jussieu, F-75005, Paris, France

(d)

Faculty of Science and Engineering, Global Center for Science and Engineering, Waseda University, 3-4-1 Okubo, Shinjuku-ku, Tokyo 169-8555, Japan.

Abstract

For a better understanding of the preservation of organic matter in clay minerals, the 1

adsorption of a model humic substance, the Gallic Acid (GA), onto a Na-montmorillonite 2

(Na-Mt) was performed in batch situation for various experimental conditions (pH=2, 5, 7) in 3

order to mimic the natural context. The adsorption efficiency and change in the clay mineral 4

were characterized via a set of complementary experimental techniques (Fourier transform 5

infrared spectroscopy, X-ray diffraction, elemental analyses). Adsorption isotherms at the 6

equilibrium were fitted with the models of Langmuir, Freundlich and Dubinin-Radushkevitch 7

allowing one to precisely quantify the adsorption through the derived fitting parameters. From 8

the adsorption data combined with complementary results of the modeled humic-clay 9

complexes, different types of interactional mechanisms were inferred as a function of 10

background acidity: (i) at pH=2 while protonated GA was the preponderant form, anionic GA 11

species can be adsorbed to the Na-Mt surface through electrostatic interaction, leading to the a 12

1

Corresponding author. E-mail address: adoum.mahamat-ahmat@bordeaux-inp.fr (A. Mahamat Ahmat).

*Revised manuscript with no changes marked

(3)

slight covering of the clay surface favoring in a second step the GA adsorption by π-π and 13

Van der Waals forces; XRD patterns corroborated via TGA and FT/IR results suggested the 14

actual intercalation of the phenolic acid within the interlayer space; (ii) At pH = 5, above the 15

pKa of phenolic acid, only 20% of the protonated form subsisted and these species were 16

adsorbed via coordinative bonding, without however any perceptible intercalation; (iii) and in 17

the regime with neutral environment (pH=7), the preponderance of GA anionic species led to 18

a poor adsorption which appeared to be only located at the external surface of the clay 19

mineral. 20

Keywords: Phenolic acids; Montmorillonite; Adsorption; Organic matter preservation.

21

(4)

1. Introduction

23

Clay minerals were recognized to stabilize soluble organic compounds through the 24

adsorption of dissolved OM in superficial horizons of soils (Gonzalez, 2002; Kögel-Knabner

25

et al., 2008; Schmidt et al., 2011; Kaiser et al., 2016). The confinement of organic compounds 26

within the interlayer space of clay minerals avoids any heterotrophic reactions and leads to 27

their preservation (Kaiser and Guggenberger, 2007; Wattel-Koekkoek and Buurman, 2004). It 28

also favors the generation of polymeric macromolecules via the condensation of single 29

monomers (Stevenson, 1982; Wang et al., 1983; Yariv and Cross, 2002). From the numerous 30

research works on the subject, and more particularly studies focusing on the understanding of 31

the interaction between clay minerals and humic substances leading to the formation of 32

humic-clay complexes, it appears that simple blocks of polymeric humic molecules or 33

elementary monomolecular compounds play a major role in the interaction with the mineral 34

surface and stabilization of the complexes (Greenland, 1971; Feng et al., 2005; Wang and

35

Xing, 2005; Chotzen et al., 2016; Chen et al., 2017). 36

The typology of physicochemical mechanisms allowing the establishment of perennial 37

association between organic compounds and mineral surfaces were extensively studied in 38

recent years. For example, ligand exchanges and cationic exchanges are reported as 39

sustainable organo-mineral interaction mechanisms (Keil and Mayer 2014; Lambert, 2018), 40

while low energy bonding such as van der Waals effects and hydrogen bonding are 41

acknowledged to establish easily reversible associations (Plante et al., 2005; Lutzow et al.,

42

2006). Aqueous media chemistry strongly constrains the preponderance of these mechanisms 43

(Arnarson and Keil, 2000). Background parameters as ionic strength and pH shape clays’ 44

state of charge as well as the degree of protonation of dissolved organic compounds. Hence, it 45

(5)

adsorption performances of humic compounds onto 2: 1 clay minerals at low pH and 47

suggested the increase of electrostatic phenomena to explain this observation. 48

Besides background properties, organic compounds intrinsic characteristics also 49

determine the extent of the adsorption and constrain the nature of preponderant bonding 50

mechanisms (Bu et al., 2017). In the presence of compounds with different spatial structures 51

and molecular functions, a competition to the occupation of adsorption sites can occur. For 52

example, the adsorption of amino acids onto montmorillonites is more effective than the 53

uptake of phenolic acids contained in the same aqueous mixture (Gao et al., 2017). 54

Understanding these mechanisms is mandatory to assess the potential use of clay minerals in 55

the adsorption of contaminants of various natures. Clays and clay minerals are investigated 56

raw, or after chemical modification, as an economically viable removal pathway of petroleum 57

by-products (Meleshyn and Tunega, 2011; Lamishane et al., 2016) and emerging 58

pharmaceutical molecules (Li et al., 2011; Thiebault et al., 2015; De Oliveira et al., 2017) 59

commonly encountered in hydrographic networks. The uptake performance may be boosted 60

via inorganic pillaring (Liu et al., 2015) or surfactant intercalation (De Oliveira and Guégan

61

2016). 62

Thus, the study of the fundamental mechanisms of organo-mineral aggregation has 63

applications in various fields. In this paper, the emphasis is placed on the organo-clay 64

association, and its further protective role. This role has been assessed in sedimentary 65

environments through different approaches (Kennedy and Wagner, 2011; Arndt and

66

Jorgensen, 2013; Mahamat Ahmat et al., 2016, 2017). It appeared that organo-clay 67

associations involving ionic exchanges and those resulting from interlayer intercalation 68

allowed efficient isolation against microbial stresses. In pedological context, the mechanism 69

has mainly been studied from the perspective of complex polymeric humic acids (Chen et al.,

70

(6)

While the main interaction mechanisms ensuring the stability of the organo-mineral 72

complexes are rather difficult to determine in the case of humic substances (Tombácz et al.,

73

2004; Chotzen et al., 2016; Chen et al., 2017), this research work aims at characterizing 74

precisely the main driving force leading to the aggregation of OM. For this purpose, we focus 75

on a simple carboxylic acid (gallic acid) as simple blocks or monomeric components in 76

polymeric humic acids. Gallic Acid (GA) is a phenolic acid found in vascularized plants, 77

inputted in soil horizons as single molecules or apart of large polyphenolic macromolecules 78

such as tannins and various ligno-cellulosic by-products. Here, in this research work, we 79

focus on the sorption mechanisms of GA onto montmorillonite under different experimental 80

conditions. 81

2. Materials and methods

82

2.1. Interaction of the carboxylic acids with a clay mineral

83

A natural Wyoming Na-montmorillonite (Na-Mt) was supplied by the Source Clay 84

Minerals Repository of the Clay Minerals Society. Its structural formula can be expressed as: 85

(Ca0.12Na0.32K0.05) [Al3.01Fe(III)0.41Mn0.01Mg0.54Ti0.02] [Si7,98Al0.02] O20(OH)4. Gallic Acid 86

(GA) was provided from Sigma Aldrich at 97.5 % purity grade and was used without further 87

treatment. This weak acid can be seen as a phenolic compound owing a carboxylic function 88

with 2 alcohol groups (Fig. 1). Its physico-chemical parameters (topological surface, pKa) are 89

summarized in Table 1. 90

The interactions between GA and Na-Mt were conducted in different experimental 91

conditions under various pH conditions: 2, 5 and 7. Typically, the adsorbent mass was 150 mg 92

in 50 mL solution spiked with a GA concentration between 10 mg. L-1 and 2 g. L-1. The pH 93

value was adjusted with both NaOH and HCl solutions (0.1 M). The solutions were stirred at 94

(7)

separated from the liquid one through a centrifugation step (5000 rpm; 10 min) and then 97

lyophilized before flash pyrolysis analyses. 98

2.2. Experimental Techniques

99

Adsorbed organic carbon was measured using flash pyrolysis (Thermo Scientific Flash 100

2000 organic analyzer) performed on organo-clay complexes in powder form. 101

The samples were also characterized via Fourier Transform Infrared Spectroscopy 102

(FT/IR) in the range 650-4000 cm-1. Measurements were realized using a Thermo Nicolet 103

6700 FT spectrometer equipped with a Deuterated Triglycine Sulfate (DTGS) detector at 104

room temperature and for different temperatures controlled by a Linkam thermal device 105

allowing us to characterize the thermal behavior on a wide range of temperature: 50-550°C. 106

The analyses were performed in transmission mode and each spectrum corresponded to the 107

average of 256 scans collected at 2 cm-1 ofresolution. 108

The d001 spacing’s of the organo-mineral complexes was determined by the first 00l 109

reflection from the X-rays patterns which were recorded on a conventional θ-θ Bragg-110

Brentano configuration by using a Thermo Electron ARL'XTRA diffractometer equipped with 111

a Cu anode (CuKα = 1.5418 Å) coupled with a Si(Li) solid detector. The diffractograms on 112

dry samples (100°C for 24 h) were performed between 1 and 24° (2θ) with an angular and 113

time steps of 0.04° and 10 s, respectively. 114

Thermal gravimetric analyses were carried out under atmospheric conditions at the 115

heating rate of 10 °C min−1 from room temperature (25°C) to 800 °C using the thermal 116

gravimetric analyzer (Model: STA PT 1600, manufactured by Linseis Company, Germany). 117

2.3. Sorption Modeling

118

The fitting of the resulting adsorption isotherms by using Langmuir, Freundlich and Dubinin– 119

Radushkevich (DR) equation models drive to numerous thermodynamic parameters allowing 120

(8)

that the whole organic molecules are adsorbed on singularized sites on the accessible surface 122

of the adsorbent, and each site hosts a unique molecule. This Langmuir model is expressed by 123

the following equation (LeVan and Vermeulen, 1981): 124

qe = qmax KL /[1 + (KL Ce )] (1)

125

where qe is adsorbed amount when equilibrium is reached (mol g-1); Ce is the

126

remaining concentration in the solution at equilibrium (mol L-1); qmax is the maximum

127

sorption capacity of the Na-Mt, and KL is the Langmuir constant (L mol-1) which is related to

128

Gibbs free energy ΔG° (kJ mol-1) through the thermodynamic equation (2): 129

ΔG° = -RTln KL (2)

130

where R represents the universal gas constant (8.314 J mol-1 K-1) and T the temperature 131

(K). 132

Freundlich and D-R equations takes into account surface heterogeneities on the 133

adsorption process and deal with the variabilities in the interaction mechanisms leading to the 134

adsorption of organic compounds that can form or be organized in multi-layers. Freundlich 135

adsorption model is a linear relation (LeVan and Vermeulen, 1981; Özcan et al., 2005) 136

expressed through the following equation: 137

ln qe = ln KF + 1/n (ln Ce) (3)

138

Where KF (g L-1) and n are constants and indicate respectively the extent of the

139

adsorption and the degree of non-linearity between GA and the smectite. Indeed, when the 140

term 1/n ranges between 0.1 and 1, it suggests that the adsorption mechanism is favorable 141

(Liu et al., 2011). D-R isotherms allow one to acquire complementary thermodynamic 142

parameters. Its equation is written as: 143

ln qe = ln qm + β*ε2 (4)

(9)

Where ε is the Polanyi potential, computed through the relation (5) 145

ε = RT ln (1+1/Ce) (5)

146

qm is the theoretical potential saturation capacity of the sorbent and β is the constant related to

147

the activity (mol2 J-2) connected to the mean free energy E of adsorption (kJ mol-1) via the 148

equation (6): 149

E = 1 / √2β (6)

150

This later parameter gives information whether the adsorption mechanism involves a cation 151

exchange or physical adsorption. Indeed, if the magnitude of E is below 8 kJ mol-1, 152

physisorption is envisaged, while for E > 8 kJ mol-1 the adsorption process follows an ion 153

exchange or a chemisorption mechanism. 154

Moreover, we used an error function (Ferror) in order to evaluate which equation model

155

was best suited to describe these processes. A lower result from the error function indicated a 156

smaller difference between adsorption capacity calculated by the model (qi cal) and the

157

experimental (qiexp). Ferror can be expressed according to the following Eq. (7)

158

Ferror = ∑ (qical – qiexp / qiexp)2 (7)

159

Where qi cal is a value of q predicted by the fitted model; qi exp is a value of q measured

160

experimentally; i indicates the values of the initial GA concentration of the experiments; and 161

P is the number of experiments performed.

162

3. Results and discussions

163

3.1. pH dependence on the adsorption of GA onto Na-Mt

164

The adsorption isotherms at the equilibrium stress out the actual affinity of GA with Na-165

(10)

when the equilibrium concentration increases. The slope of this growth attenuates at high 167

starting concentrations emphasizing a saturation state, excepted for the isotherm realized at 168

pH=5. 169

GA adsorption isotherms are properly fitted by the three equation models used as r2 values 170

display values between 0.95 and 0.99 and Ferror values are between 0.0010 and 0.130 for GA

171

(Table 2). Based on r2 values, experimental data seem to be better adjusted to the Langmuir 172

model, however its function errors are higher than 0.1 and to those for both Freundlich and 173

DR equations, which spread out from 0.001 to 0.004. This is a side effect of the logarithmic 174

scale adopted in Freundlich and DR representations. Although Langmuir equation properly 175

fitted experimental data, the two latter equations appear to be more suitable for modeling the 176

adsorption of the phenolic acid onto the clay mineral surface. Indeed, clay mineral shows a 177

heterogeneous surface leading to a distribution of several adsorption sites that are taken into 178

account in both Freundlich and DR equation models.

179

Under low pH conditions (i.e. below the pKa of GA), the adsorption is particularly enhanced, 180

with the predominance of the protonated form of GA (i.e. acidic form) in solution, due to the 181

decrease of the repulsion between the neutral GA and Na-Mt. However, the situation may be 182

rather twisted with antagonist effects. Indeed, under such pH conditions (pH=2), the pH value 183

is lower than the pH of zero net proton charge (pHZNPC) of Na-Mt, estimated about 4.5 184

(Tombàcz and Szekeres, 2006).For pH < pHZNPC, the electric charge of the edge-sites of clay 185

mineral surface changes from anionic to cationic, favoring the adsorption of phenolic acids in 186

their anionic form, (i.e. base) allowing a possible ion exchange process which enhances the 187

amount of adsorbed organic acids. In contrast, at pH=5 and 7, since both the clay mineral 188

edge-sites and the equilibrium ratio between the acid/base forms change (Fig. 3), the 189

(11)

adsorption isotherms while reducing the adsorbed amounts. This lowering is due to the 191

repulsion between anionic charges of both GA and Na-Mt. 192

193

Fig. 4 shows the X-ray diffraction patterns evolution of the GA-clay complexes (dried 194

for 48 hours at 100°C to prevent any water molecules in the interlayer space that may 195

interfere in the interpretation and understanding of a possible intercalation of the GA) 196

following the starting GA concentration. The diffraction patterns display several diffraction 197

peaks located between 5° and 9° (2θ) related to the 00l reflections. At pH=2, the 00l reflection 198

shifts to lower angle values, attesting the effective intercalation of the phenolic compound. As 199

well, it is interesting to remark the narrowing of these reflections suggesting an enhancement 200

of the organization in the layered material probably due to the increase of the GA density 201

within the clay–GA complexes. The basal spacing of GA-Mt complexes (i.e. d001), estimated 202

with the angular position of the 00l reflection, increases from 9.6 Å for a dehydrated clay 203

mineral to about 13.3 Å at pH=2. While an intercalation occurs at low pH, this phenomenon is 204

not observed for the other two pH, where a slight increase of the interlayer space to a value of 205

11 Å can be noticed (Fig. 5). This is probably due to both the clay mineral charge density and 206

the preponderance of the neutral (RCO2H) form which play an important role in the 207

adsorption as well its associated interaction mechanism (e.g. coordinative bonding ). 208

The contribution of FT/IR spectroscopy gives important information in the 209

characterization of GA-clay complexes derived from GA and Mt and confirms its actual 210

adsorption. Absorption bands observed at 1350, 1384 and 1470 cm-1 are assigned to stretching 211

modes of C – O. and bending modes of C-H of O-H respectively (Fig. 6). The strong vibration 212

at 1700 cm-1 was assigned to C=O stretching of GA’s carboxylic function. Besides the 213

existence of a drift or a temperature gradient in temperature during the FT/IR experiments, the 214

(12)

GA which is decomposed at about 300°C estimated through thermal gravimetric 216

measurements (Fig. 8). This last observation about the preservation of GA at high temperature 217

is related to its confinement within the interlayer space of Mt obtained for GA at pH=2. 218

TG analyses allow assessing the loss of weight of the adsorbent during a gradual 219

heating. Organo-clay minerals display usually three main weight losses during the heating. 220

The first one is associated to the evaporation of free and adsorbed water, ranging between the 221

initial temperature and 150°C. The second one is related to the thermal oxidation of adsorbed 222

organic compounds between 150 and 600°C with a maximal decomposition temperature 223

related to the characteristics of organic moieties. Finally, for temperatures higher than 600°C, 224

only the dehydroxylation of clay minerals is expected (Xie et al., 2001). Fig. 8 gathers TG and 225

DTG curves of GA-Na-Mt composites. Weight losses associated to the dehydration remains 226

independent from pH values. Corresponding DTG peaks occur between 98 and 105°C. With 227

growing temperature, the distinction between each sample is more pronounced. Hence, 228

between 150 and 600°C, no significant weight loss is displayed at pH=7, whereas noteworthy 229

weight losses are observed at the lower pH values (2 and 5). The first decomposition 230

temperature is noticed at 247°C and corresponds to the decomposition of acidic group of GA 231

(Rao et al., 1981; Hussein et al., 2009). In the organic matter combustion range, GA-Na-Mt 232

complexes aggregated at pH = 5 display one loss of weight (i.e. 340°C), while those formed at 233

pH=2 exhibit two noticeable losses at 310 and 410°C (Fig. 8). This splitting after interaction 234

at pH=2 might be explained by the distinction between GA adsorbed onto the external surface 235

of the clay mineral (i.e. lower temperature) and intercalated GA (i.e. higher temperature) (Zhu

236

et al., 2017). Hence, this result appears to be consistent with XRD results, in which an 237

intercalation of GA only occurs after interaction at pH=2. 238

3.2. Adsorption mechanisms and geochemical model of GA-Mt complexes

(13)

The observation that GA adsorption is higher at pH < pKa is in agreement with the 240

results of Rabiei et al., 2016 which stressed out higher GA adsorptions at pH=3 (33%) 241

compared to their experiments performed at pH=7 (20%). This pH-dependency is related to 242

the protonation capacity of GA’s polar appendage (COOH / COO-). It must be noted however, 243

that the pH control of GA adsorption cannot be generalized to the interaction of other phenol-244

based molecules. The study by Dolaksis et al., 2018 for instance, suggested that the 245

adsorption of phenolic cores devoid of polarizing COOH appendix is pH–independent. They 246

showed that low energy mechanisms (hydrogen bridges) drive the adsorption of 247

chlorophenols and nitrophenols onto silicate surfaces. 248

At a pH < pKa, GA is mainly protonated (neutral) and this form represents about 99.6 % of 249

the chemical species at pH=2 (Fig. 3). However, in acidic condition, the charge density of the 250

edge-sites of Na-Mt changes and switches below the pH of zero net proton charge estimated 251

at 4.5 (Table 1). Here, despite the possibility of an alteration of the structure and probably the 252

chemical composition of the layered material, the adsorption of GA appears to be enhanced in 253

acidic conditions. The parameters derived from the fitting procedure, and more particularly 254

those of the D-R model giving a free energy of adsorption E slightly above 8 kJ mol-1, 255

underlining the possibility of chemical process for adsorption or at least strong electrostatic 256

interactions. Although being less known and implied in the adsorption of anionic compounds 257

onto Mt, electrostatic interaction was recognized as the main driving force leading to the 258

intercalation of various kind of anionic compounds: tannic and benzoic acids, anionic 259

surfactants (Yan et al., 2007; Zhang et al., 2012; An and Dultz 2007), that may occur 260

nevertheless, in particular experimental conditions (low pH range) as it is the case here. 261

Moreover, the increase of the basal spacing of Na-Mt after the interaction with GA at pH=2 262

(i.e. + 3.7 Å) is consistent with the molecular size of GA (i.e. z=3.7 Å, Figure 1), emphasizing 263

(14)

The adsorption of anionic species, even weak, acts as a coating that may favor further 265

adsorption of organic molecules as organoclay materials do, nevertheless cannot only explain 266

the totality of the adsorbed amount reaching about 1.5 x 10-4 mol g-1. Indeed, GA is in anionic 267

form at a low concentration. The amount of anionic species may increase during the sorption 268

mechanism since the uptake involves the displacement of acid basic equilibrium. This may 269

allow in fine the adsorption of higher amount of anionic species. However, the effect of this 270

acid basic shift is limited. Protonated species of GA remain preponderant and are likely to be 271

adsorbed through other bonding mechanisms. Since GA is mainly neutral at pH < pKa as 272

explained before, its adsorption should be driven through physisorption mechanisms such as 273

molecular interaction (π- π interaction, van der Waals forces) with the prior adsorbed 274

molecules and by coordinative bonding through inorganic exchangeable cations located 275

within the interlayer space and with the carboxylic moieties as both FT / IR and XRD data 276

highlighted. A recent work stressed out the importance of ion-dipole interaction (a 277

coordinative bonding mechanism) as the main driving force for the adsorption of nonionic 278

surfactants (Guégan et al., 2017) onto a Mt surface and can be according to previous studies 279

in the literature (Sonon and Thompson, 2005; Deng et al., 2006) extended to nonionic 280

compounds and here GA in its neutral form (Fig. 7). 281

The preponderance and the role of this interaction is confirmed at a pH value of 5 where 282

the maximum adsorbed amounts at the equilibrium reaches 1.3 x 10-4 mol g-1. Based on the 283

diagram of preponderance of GA species in regards to pH, its neutral form represents about 284

20% at a pH=5 (Fig. 3). The experiment with the highest starting concentration of 2 g L-1 285

leads to 5.88 x 10-4 mol g-1 (if one does the hypothesis that such amount is adsorbed – number 286

of GA moles normalized to the mass of Mt introduced) where 20% are in the RCO2H form 287

(neutral one), thus driving to a value of 1.176 x 10-4 mol g-1, close to the experimental one 288

(15)

capacity as well as an anionic charge surface increasing the repulsion between adsorbent and 290

organic anions. Hence, the protonated form of GA is favored for adsorption, and may interact 291

to the inorganic exchangeable cations through coordinative bonding forces. However, it is 292

important to mention that 20% of the inorganic exchangeable cations are located onto the 293

external surface of the clay mineral (Swartzen-Allen and Matijević, 1975; Shainberg et al.,

294

1980; Patzko and Decany, 1993; Logdson and Laird, 2004), where they can be easily 295

mobilized for a cation-exchange or other interactions such as coordinative bonding or other 296

mechanism such as cationic bridges involving the anionic species and divalent compensating 297

cations. Here, the adsorption of GA does not lead to any intercalation as the XRD data 298

displayed and exclusively remains on the external surface. The presence of the inorganic 299

exchangeable cations at about 20% of the CEC on the external surface can be easily mobilized 300

through coordinative bonding or via complexation reactions (Fig. 7) with the phenolic acid of 301

which adsorbed amount match a value lesser than the 20% of the CEC (1.6 x 10-4 mol g-1). 302

Similar observations are noticed at pH=7, where the adsorbed amount of neutral GA is not 303

enough to lead to any intercalation as both XRD (Fig. 5) and FT/ IR data showed and restrict 304

at this pH range principally anionic species in the adsorption. Indeed, in contrast to the 305

previous pH, where the proportion of neutral species of GA represents about 96 and 20 % at 306

pH=2 and 5 respectively, at a pH=7, it is insignificant with only 0.25 % (Fig. 3). While, the 307

anionic form of GA is preponderant, an adsorption is surprisingly observed without however 308

any intercalation as it was the case for pH=5, leading to an adsorbed amount of about 1 x 10-4 309

mol g-1 (Fig. 2). Here, the adsorption may involve interaction mechanisms such as cationic 310

bridges between anionic GA and divalent inorganic cations. Hence, even mostly compensated 311

with Na+, around 20% of the compensating inorganic cations of Na-Mt are Ca2+, that are able 312

to sorb anionic species through cationic bridges (Errais et al., 2012; Thiebault et al., 2016;

313

(16)

pH = 5, although the distinction between these mechanisms is not possible based on these 315

experiments (Fig. 7). 316

3.3 Putting into perspective the behavior of GA with other phenolic-based compounds

317

and clay minerals

318

The adsorption of GA is enhanced at low pH, as polyphenolic humic molecules behave 319

(Feng et al., 2005; Zhang et al., 2012; Chotzen et al., 2016). Polyphenolic molecules as fulvic 320

and humic acids, exhibit indeed greater adsorption rates with growing acidity (Gouré-Douby

321

et al., 2018). For instance, the study of Chen et al. (2017) focusing on the adsorption of soil 322

macromolecular humic acids onto both montmorillonite and kaolinite, pointed out the 323

enhancement of the adsorption at low pH. GA adopts a similar behavior when background pH 324

ranges below its pKa and allows a preeminence of its protonated species. 325

In the case of natural polyphenolic molecules (humic substances), it has been repeatedly 326

observed that in addition to background acidity, the nature of the adsorbent has a predominant 327

role. Clay minerals of the 1:1 group such as kaolinite appear more conducive to the adsorption 328

of humic –type of polyphenols. This is promoted by edges electrostatic phenomena. The 329

extent of the adsorption through edge electrostatic interaction principally depends on the 330

physico-chemical properties and mineralogy of a clay mineral, and more particularly its pH of 331

zero net proton charge (pH ZNPC) which is estimated to 4.5 for Na-Mt (Tombàcz and Szekeres,

332

2006), lower than other soil clayey components such as kaolinite (Wiliams and Wiliams,

333

1978; Gupta and Miller, 2010). With its particular properties: a pHZNPC and background 334

protonation leading to a proper dispersion of clay mineral particles, kaolinite favors the 335

adsorption of macromolecular humic acids at large content in contrast to Na-Mt besides its 336

large CEC and specific surface area values (Chotzen et al., 2016). With similar background 337

(17)

be added to the key criteria ruling GA’s adsorption However, care should be taken in 340

comparing the uptake of these phenolic-based compounds as differences in molecular masses, 341

spatial arrangements and number of charges may induce differences in sorptive behaviors. 342

Although being poorer from quantitative perspective, the interaction of GA with Na-Mt 343

remains interesting from a protective point of view since its neutral form may intercalate (Fig.

344

2; Fig. 7) under specific acidic conditions (pH < pKa). Interlayers isolations are reported to 345

reduce bioavailability (Theng et al., 2001) and help to prevent biotic redox transformations. 346

In our case, the pH dependence of the adsorption appears to be consistent with the 347

trends shown by other phenolic and polyphenolic molecules. However, the nature of the 348

interactional mechanisms varies according to the molecules and does not display any 349

systematic pattern. Although our results suggest intercalation and surface complexations via 350

weak mechanisms (coordinative bonding), ligand exchanges are often involved in the 351

interaction of long phenolic chains. 352

This encourages to complexify our view of the aggregation modalities of clays and 353

phenol-like molecules and the possibility of organic matter stabilization process that it 354

inducts. Na-Mt interactional mechanisms seems to differ whether the involved organic 355

molecule are heavy macromolecular acids or singularized phenolic compounds. Thus, in a 356

pedological environment where micro-fauna regime induces high rates of polymeric lyses and 357

solubilizes a high quantity of phenolic acids, the stabilization process follows a different 358

pathway from the configuration where polymeric humic acids are weakly degraded. In the 359

first case, our data suggest that intercalation should prevail during the adsorption of poly-360

phenolic acids by-products in acidic pH conditions, while several studies point out adsorption 361

on the edge and ligand exchanges in the ecological context where polymeric forms are 362

(18)

4. Conclusions

364

The series of organo-clay interactions performed here under evolutionary batch 365

equilibrium conditions allowed to characterize the adsorption mechanisms of GA, a common 366

phenolic acid in natural pedological media, onto Na-Mt. The parameters derived from fitting 367

of GA data with a reasonable agreement (high r2 values) to the adsorption models used: 368

Langmuir, Freundlich and D-R equations, gave pertinent insights for a proper description of 369

the phenomena under different approaches which pointed out the good affinity of the phenolic 370

acid to Na-Mt. Additional experimental results obtained by FT/IR, TGA and XRD 371

corroborated the actual adsorption of GA onto Na-Mt. 372

From the set of data, it appears that this monomolecular acid is mainly adsorbed 373

through coordinative bonding interaction in contrast to macromolecular humic acids at low 374

pH range, where ligand exchanges or complexation reaction occur according to the literature. 375

Here, GA, a singularized element of large pedological compounds, interacts with the clayey 376

mineral mainly via both surface and interlayer processes at low pH, leading to its confinement 377

within the interlayer space which may allow a sustainable preservation of the organic matter 378

in that way. 379

Acknowledgement

380

This study was supported by the project MONITOPOL funded by the French region 381

Centre Val de Loire (grant number 00117247). The authors are grateful to Marielle Hatton for 382

her analytical contribution. 383

Figures Captions

384

(19)

Fig. 2: Adsorption isotherms of Gallic Acid. Beige squares represent the data obtained at pH 386

2, green ones are for the data measured at pH 5 and the red triangles represent those collected 387

at pH 5. The continuous line represents Langmuir model fit. 388

Fig. 3: GA speciation following pH conditions. 389

Fig. 4: Graph of the 3D evolution of XRD diffraction patterns of dehydrated Gallic Acid (GA) 390

and Na-Mt composite samples as a function of the starting GA concentrations in solution for 391

pH=2. Only the results going up to 0.1 g L-1 arepresented here, since no particular evolution 392

of (001) planes was observed beyond this concentration. 393

Fig. 5: Evolution of the d001 basal spacing determined by the 00l reflection of Na-Mt layers 394

obtained for GA / Na-Mt humic-clay like samples. 395

Fig. 6: FT/IR spectra of GA/Na-Mt composite samples (pH=2): thermic evolution. 396

Fig.7: Schematic representation of the possible adsorption mechanisms leading to the 397

intercalation of GA within the interlayer of a Na-Mt at pH=2, and adsorption for the pH=5 398

and 7. 399

Fig. 8: TG (solid lines) and DTG (dashed lines) curves of GA/Na-Mt (CGA = 0.1 g L-1) 400

composite samples after interaction at pH=2 (dark gray lines); pH=5 (dark lines) and pH=7 401

(light gray lines). 402

Tables Captions

403

Table 1: Physico-chemical properties of implemented clay mineral (Na-Mt) and GA. 404

Table 2: Adsorption isotherm constants determined with Langmuir, Freundlich, and Dubinin-405

Radushkevich model fit for the adsorption of GA onto Na-Mt for different pH. 406

Supplementary data

(20)

Appendix 1: Spectroscopic observations (FT/IR) of GA-Na-Mt complexes aggregated at 408

different levels of GA starting concentrations. 409

References

410

An, J-H., Dultz, S., 2007. Adsorption of tannic acid on chitosan-montmorillonite as a function 411

of pH and surface charge properties. Applied Clay Science 36, 256-264. 412

Arnarson, T.S., Keil, R.G., 2000. Mechanisms of pore water organic matter adsorption to 413

montmorillonite. Marine Chemistry 71, 309-320. 414

Arndt, S., Jorgensen, B.B., 2013. Quantifying the degradation of organic matter in marine 415

sediments: A review and synthesis. Earth-Science Reviews 123, 53-86. 416

Bu, H., Yuan, P., Liu, H., Liu, D., Liu, J., He, H., Li, Z. (2017). Effects of complexation 417

between organic matter (OM) and clay mineral on OM pyrolysis. Geochimica et 418

Cosmochimica Acta, 212, 1-15. 419

Chen, H., Koopal, L.K., Xiong, L.K., Xiong, J., Avena, M., Tan, W., 2017. Mechanisms of 420

soil humic acid adsorption onto montmorillonite and kaolinite. Journal of Colloid and 421

Interface Science 504, 457-467. 422

Chotzen, A.R., Polubesova, T., Chefetz, B., Mishael, Y.G., 2016. Adsorption of soil-derived 423

humic acid by seven clay minerals, a systematic study. Clays and Clay Minerals 64, 424

628-638. 425

De Oliveira, T., Guegan, R., 2016. Coupled Organoclay / Micelle Action for the Adsorption 426

of Diclofenac. Environmental Science & Technology 50, 10209-10215 427

De Oliveira, T., Guégan, R., Thiebault, T., Le Milbeau C., Muller F., Teixeira, V., Giovanela, 428

M., Boussafir, M., 2017. Adsorption of Diclofenac onto organoclays: effect of 429

surfactant and environmental (pH and temperature) conditions. Journal of Hazardous 430

(21)

Deng, Y., Dixon, J.B., Norman White, G., 2006. Bonding mechanisms and conformation of 432

poly(ethylene oxide)-based surfactants in interlayer of smectite. Colloid and Polymer 433

Science 284, 347-356. 434

Derakhshani, E., Naghizadeh, A., 2018. Optimization of humic acid removal by adsorption 435

onto bentonite and montmorillonite nanoparticles. Journal of Molecular Liquids 259, 436

76-81. 437

Dolaksiz, E.Y., Temel, F., Tabakci, M., 2018. Adsorption of phenolic compounds onto 438

calix[4]arene-bonded silica gels from aqueous solutions. Reactive and Functional 439

Polymers 126, 27 - 35. 440

Errais, E., Duplay, J., Elhabiri,M., Khodja, M., Ocampo, R., Baltenweck-Guyot, R., 441

Darragi,F., 2012. Anionic RR120 dye adsorption onto raw clay: Surface properties and 442

adsorption mechanism. Colloids and Surfaces A: Physicochemical and Engineering 443

Aspects 403, 69 - 78. 444

Feng, X., Simpson, A.J., Simpson, M.J., 2005. Chemical and mineralogical controls on humic 445

acid sorption to clay mineral surfaces. Organic Geochemistry 36, 1553-1566. 446

Gao, J., Jansen, B., Cerli, B., Helmus, R., Mikutta, R., Dultz, S., Guggenberger, G., Kalbitz, 447

K., 2017. Competition and surface conditioning alter the adsorption of phenolic and 448

amino acids on soil minerals. European Journal of Soil Science 68, 667-677. 449

Gonzalez, J- M., 2002. Role of clay minerals on soil organic matter and humification. Ph.D 450

thesis of Iowa State University, 212 pp. 451

Gouré-Douby, H., Martias, C., Smith, A., Villandier, N., Sol, V., Gloaguen, V., Feuillade, G., 452

2018. Adsorption of fulvic and humic like acids on surfaces of clays: Relation with 453

(22)

Guégan, R., Veron, E., Le Forestier, L., Ogawa, M., Cadars, S., 2017 Structure and dynamics 455

of nonionic surfactant aggregates in layered materials. Langmuir 33, 9759-9771. 456

Gupta, V., Miller, J.D., 2010. Surface force measurements at the basal planes of ordered 457

kaolinite particles. Journal of Colloid and Interface Science 344, 362-371. 458

Greenland, D.J., 1971. Interaction between humic and fulvic acids and clays. Soil Science, 459

111 34-41. 460

Hussein, M.Z.B., Ghotbi, M.Y., Yahaya, A.H., Abd Rahman, M.Z., 2009. Synthesis and 461

characterization of (zinc-layered-gallate) nanohybrid using structural memory effect. 462

Materials Chemistry and Physics 113, 491-496. 463

Kaiser, K., Guggenberger, G., 2007. Sorptive stabilization of organic matter by microporous 464

goethite: Sorption into small pores vs. surface complexation. European Journal of Soil 465

Science 58, 45-59. 466

Kaiser, K., Zederer, D.P., Ellerbrock, R.H., Sommer, M., Ludwig, B., 2016. Effects of 467

mineral characteristics on content, composition, and stability of organic matter fractions 468

separated from seven forest topsoils of different pedogenesis. Geoderma 263, 1-7. 469

Keil, R.G., Mayer, L.M., 2014. Mineral matrices and organic matter. Treatise on 470

Geochemistry (Second Edition) 12, 337-359. 471

Kennedy, M.J., Wagner, T., 2011. Clay mineral continental amplifier for marine carbon 472

sequestration in a greenhouse ocean. Proceedings of the National Academy of Sciences 473

of the United States of America 108, 9776-9781. 474

Kögel-Knabner, I., Guggenberger, G., Kleber, M., 2008. Organo-mineral associations in 475

temperate soils: Integrating biology, mineralogy, and organic matter chemistry. Journal 476

of Plant Nutrition and Soil Science-Zeitschrift für Pflanzenernahrung und Bodenkunde 477

(23)

Lambert, J.F., 2018. Organic pollutants adsorption on clay minerals. Developments in Clay 479

Science 9, 195-253. 480

Lamishhane, S., Bal Krishna;K.C., Sarukkalige, R., 2016. Polycylic Aromatic Hyrdrocarbons 481

(PAHs) removal by sorption: A review. Chemosphere 148, 336-353. 482

LeVan, M.D., Vermeulen,T., 1981. Binary Langmuir and Freundlich isotherms for ideal 483

adsorbed solutions. The Journal of Physical Chemistry 85, 3247-3250. 484

Li, Z., Chang, P-H., Jiang. W-T., Jean, J-S., Hong, H., Liao, L., 2011. Removal of 485

diphenhydramine from water by swelling clay minerals. Journal of Colloid and 486

Interface Science 360, 227-232. 487

Liu, R., Zhang, B., Mei, D., Zhang, H., Liu, J., 2011. Adsorption of methyl violet from 488

aqueous solution by halloysite nanotubes. Desalination 268, 111-116. 489

Liu, Y., Dong, C., Wei, W., Yuan, W., Li, K., 2015. Adsorption of levofloxacin onto an iron-490

pillared montmorillonite (clay mineral): Kinetics, equilibrium and mechanism. Applied 491

Clay Science 118, 301-307. 492

Logsdon, S.D., Laird, D.A., 2004. Electrical conductivity spectra spectra of smectites as 493

influenced by saturating cation and humidity. Clays and Clay Minerals 52, 411 – 420. 494

Lutzow, M., Kögel-Knabner, I., Ekschmitt, K., Matzner, E., Guggenberger, G., Marschner, 495

B., Flessa, H., 2006. Stabilization of organic matter in temperate soils: Mechanisms and 496

their relevance under different soil conditions-A review. European Journal of Soil 497

Science 57, 426-445. 498

Mahamat-Ahmat, A., Boussafir, M., Le Milbeau, C., Guégan, R., Valdès, J, Guiñez, M., 499

Sifeddine, A., Le Forestier, L., 2016. Organic matter-clay interaction along a seawater 500

column of the Eastern Pacific upwelling system (Antofagasta bay, Chile): implications 501

(24)

Mahamat-Ahmat, A., Boussafir, M., Le Milbeau, Guégan, R., De Oliveira, T., Le Forestier, 503

L., 2017. Organic matter and clay interaction in a meromictic lake: implications for 504

source rock OM preservation (Lac Pavin, Puy-de-Dôme, France). Organic 505

Geochemistry 109, 47-57. 506

Meleshyn, A., Tunega., 2011. Adsorption of phenanthrene on Na-Montmorillonite: A model 507

study. Geoderma 169, 41-46. 508

Nam Jang, B., Wilkie, C.A., 2004. A TGA/FTIR and mass spectral study on the thermal 509

degradation of bisphenol A polycarbonate. Polymer Degradation and Stability 86, 419-510

430. 511

Özcan, A.S., Erdem, B., Özcan, A., 2005. Adsorption of acid blue 193 from aqueous solutions 512

onto BTMA-bentonite. Colloids and Surfaces A: Physicochemical and Engineering 513

Aspects 266, 73-81. 514

Patzko., A., Decany., I., 1993 Ion exchange and molecular adsorption of a cationic surfactant 515

on clay minerals . Colloids and Surfaces A: Physicochemical and Engineering Aspects. 516

71 (1993) 299-307. 517

Plante, A.F., Pernes, M., Chenu, C., 2005. Changes in clay-associated organic matter quality 518

in a C depletion sequence as measured by differential thermal analyses. Geoderma 129, 519

186-199. 520

Rabiei, M., Sabahi, H., Rezayan, A.H., 2016. Gallic acid-loaded montmorillonite 521

nanostructure as a new controlled release system. Applied Clay Science 119, 236-242. 522

Rao, C.R.M., Reddy, V.B., Mehrotra, P.N., 1981. Derivatographic studies of gallic acid. 523

Thermochimica Acta 46, 65-69.Shainberg, I., Oster, J.D., Wood, J.D., 1980. Sodium / 524

Calcium Exchange in Montmorillonite and Illite Suspensions 1, Soil Science Society of 525

(25)

Swartzen-Allen, S.L., Matijević, E., 1975. Colloid and surface properties of clay suspensions: 527

II. Electrophoresis and cation adsorption of montmorillonite. Journal of Colloid 528

Interface Science 50, 143–153. 529

Sonon, L.S., Thompson, M.L., 2005. Sorption of a nonionic polyoxyethylene lauryl ether 530

surfactant by 2:1 layer silicates. Clays and Clay minerals 53, 45-54. 531

Schmidt, M.W.I., Torn, M.S., Abiven, S., Dittmar, T., Guggenberger, G., Janssens, I.A., 532

Kleber, M., Kögel-Knabner, I., Lehmann, J., Manning, D.A.C., Nannipieri, P., Rasse, 533

D.P., Weiner, S., Trumbore, S.E., 2011. Persistence of soil organic matter as an 534

ecosystem property. Nature 478, 49-56. 535

Stevenson, F.J. 1982. Humus Chemistry. John Wiley & Sons, New York. 536

Theng, B.K.G., Aslabie, J., Fraser, R., 2001. Bioavailability of phenanthrene intercalated into 537

an alkylamonium-montmorillonite clay. Soil Biology and Biogeochemistry 33, 845-848. 538

Thiebault, T., Guégan, R., Boussafir, M., 2015. Adsorption mechanisms of emerging micro-539

pollutants with a clay mineral: case of tramadol and doxepine pharmaceutical products. 540

Journal of Colloid and Interface Science 453, 1-8. 541

Thiebault, T., Boussafir, M., Guégan, R., Le Milbeau, C., Le Forestier, L., 2016. Clayey-sand 542

filter for the removal of pharmaceuticals from wastewater effluent: percolation 543

experiments. Environmental Science Water Research & Technology 2, 529-538. 544

Tombácz, E., Libor, Z., Illés, E., Majzik, A., Klumpp, E., 2004. The role of reactive surface 545

sites and complexation by humic acids in the interaction of clay mineral and iron oxide 546

particles. Organic Geochemistry 35, 257-267. 547

Tombácz, E., Szekeres, M., 2006. Surface charge heterogeneity of kaolinite in aqueous 548

(26)

Wang, K., Xing., B., 2005. Structural and sorption characteristics of adsorbed humic acid on 550

clay minerals. Journal of Environmental Quality, 34 342-349. 551

Wang, T.S.C., Wang, M.C., Ferng, Y.L., 1983. Catalytic synthesis of humic substances by 552

natural clays, silts, and soils. Soil Science 135, 350-360. 553

Wattel-Koekkoek, E.J.W., Buurman, P., 2004. Mean residence time of kaolinite and smectite-554

bound organic matter in Mozambiquan soils. Soil Science Society of America Journal 555

68, 154-161. 556

Williams, D.J.A., Williams, K.P., 1978. Electrophoresis and zeta potential of kaolinite. 557

Journal of Colloid and Interface Science 65, 79-87. 558

Xie, W., Gao, Z., Liu, K., Pan, W.-P., Vaia, R., Hunter, D., Singh, A., 2001. Thermal 559

characterization of organically modified montmorillonite. Thermochimica Acta 367-560

368, 339-350. 561

Yan, L., Wang, J., Yu, H., Wei, Q., Du, B., Shan, X., 2007. Adsorption of benzoic acid by 562

CTAB exchanged montmorillonite. Applied Clay Science 37, 226-230. 563

Yariv, S., Cross, H., 2002. Organo-clay complexes and interactions. Marcel Dekker, New 564

York 688pp. 565

Zhang, L., Luo, L., Zhang, S., 2012. Integrated investigations on the adsorption mechanisms 566

of fulvic and humic acids on three clay minerals. Colloids and Surfaces A: 567

Physicochemical and Engineering Aspects 406, 84-90. 568

Zhu, J., Zhang, P., Qing, Y., Wen, K., Su, X., Ma, L., Wei, J., Liu, H., He, H., Xi, Y., 2017. 569

Novel intercalation mechanism of zwitterionic surfactant modified montmorillonites. 570

Applied Clay Science 141, 265- 271. 571

572

(27)

Highlights

- Gallic acid (GA) adsorption onto Na-Mt was performed in batch conditions; - Data were fitted to Dubinin-Radushkevitch, Langmuir and Freundlich equations; - Following the acidity, both edge and interlayer interactions modes were observed; - GA intercalates under its neutral form;

(28)

Properties Values

Na-MMt C.E.C (mol.g-1) 8.10-4

pznc 4.5

GA COOH pKa 4.4

Surface area (Å2) 98

Table 1

(29)
(30)

x=7.64 Å

y=5.87 Å

z=3.72 Å

Figure 1

(31)
(32)
(33)

Figure 4

(34)
(35)

Figure 6

(36)

Na

+

Cla

y Mineral Na-Mt

d

001

≈9.7Å

In

tercalation of GA

pH < pKa

d

001

≈13.7Å

External adsorption of GA

pH > pKa

d

001

≈9.7Å

(37)

0 200 400 600 800 1000

T

em

p

era

ture

C

)

0 -0.01 -0.02 -0.03 95 97 96 98 99 100



weight loss

(mg.°C

-1

)

W

eight

(%)

pH 2 pH 5 pH 7 105°C 247°C 310-410°C

Références

Documents relatifs

This Environmental Health Criteria (EHC) monograph deals with the health hazards associated with bentonite, kaolin, and common clay, which are commercially important clay products,

The following methodological chain is based on (i) spectral preprocessing to transform reflectance spectra in a standardized form and (ii) linear and nonlinear unmixing algorithms

Surface oxygen (Os) DPs are in red, water oxygen (Ow) in light blue, and sodium cations in purple; for carbamazepine (left), the snitrogen are in blue and the oxygen is in green;

The chemical modification of clay mineral via the use of cationic surfactants, like the benzyl dimethyl tetradecyl ammonium (BDTA), for the preparation of

For tramadol and doxepin drugs defined as emerging micro-pollutants due to their presence at significant concentration in numerous water compartments, the

Thus, this observation demonstrates that the whole surface (i.e. internal and external) of clay mineral is accessible for the adsorption leading to multiple

The electrochemical oxidation of calcium channel blocker drug of Nimodipine at a sodium montmorillonite modified GCE was identified in buffered solutions and a

Extensive analyses have been conducted on various hand-collected and bee- collected pollens but little information is available on the chemical changes occuring in pollen as